Abstract
Chemical modification and spontaneous loss of nucleotide bases from DNA are estimated to occur at the rate of thousands per human cell per day. DNA base excision repair (BER) is a critical mechanism for repairing such lesions in nuclear and mitochondrial DNA. Defective expression or function of proteins required for BER or proteins that regulate BER have been consistently associated with neurological dysfunction and disease in humans. Recent studies suggest that DNA lesions in the nuclear and mitochondrial compartments and the cellular response to those lesions have a profound effect on cellular energy homeostasis, mitochondrial function and cellular bioenergetics, with especially strong influence on neurological function. Further studies in this area could lead to novel approaches to prevent and treat human neurodegenerative disease.
Keywords: Base excision repair, oxidative damage, mitochondrial DNA, neurodegeneration, PARP-1
INTRODUCTION
DNA and other nucleic acids are susceptible to damage from endogenous and exogenous sources, including oxygen free radicals, UV-light, high energy cosmic particles and chemical compounds. Cells have mechanisms to prevent or limit the adverse consequences of such damage, when it occurs, including specialized DNA repair pathways.
The DNA base excision repair (BER) pathway repairs the majority of endogenous DNA damage. BER plays an important role in the brain homeostasis at least for three reasons. First, oxidative DNA damage is a frequent event in brain cells, because brain consumes more oxygen and has higher metabolic rate and lower antioxidant capacity than many other tissues [1]. Second, BER is critical for maintaining the integrity of the mitochondrial genome, which is essential for neurological function [2]. Third, BER repairs DNA damage caused by physiological neuronal activity [3, 4].
DNA base excision repair (BER)
BER is the primary mechanism for repair of DNA base damage, AP sites and variants of single-strand breaks. BER begins with the identification and removal of a damaged base by a DNA glycosylase leaving an AP-site. The AP-site is then incised on the 5′-side by AP endonuclease 1 (APE-1), followed by gap filling with a DNA polymerase and finally joining of the DNA ends by a DNA ligase (Fig. 1). In principle, BER can be reconstituted in vitro with four proteins; a DNA glycosyase, APE-1, DNA polymerase β (Polβ), and DNA ligase III (Lig III). In the cells, however, BER involves a number of other proteins; some are essential such as APE-1 and Polβ, while others play a more regulatory role and coordinate the BER response to damage such as X-ray repair cross-complementing protein 1 (XRCC1) and proliferative nuclear antigen (PCNA). BER is also regulated by the cell cycle, protein-protein interactions, and post-translational modifications [5, 6]. Some BER proteins also play roles in the adaptive immune response and in RNA quality control[7–10], indicating a multifunctional role in nucleic acid metabolism(Fig. 2).
Figure 1.
DNA Base Excision Repair Pathways. Base excision repair (BER) can begin with the recognition of a damaged base by a DNA glycosylase, an AP site by APE1, or a single-strand break by end-processing proteins. BER can proceed via incorporation of a single nucleotide (short patch, SP), or by incorporation of two or several nucleotides (long patch, LP,). Red rings around proteins denote that they are lethal if knocked out in mice. Black ring around multiple glycosylases denote that these proteins are non-lethal when knocked-out in mice. Mutations in TDP1, APTX and PNKP are associated with the following human diseases: Spinocerebellar ataxia with axonal neuropathy (SCAN1), Ataxia with oculomotor apraxia 1 (AOA1) and Microcephaly, seizures and developmental delay (MCSZ) and are noted on the figure.
Figure 2.

Base Excision Repair is a Central Player. The proteins involved in base excision repair (BER) are denoted. Arrows indicate significant BER protein or pathway associations with other pathways, features or endpoints.
BER enzymes act on both the nuclear and the mitochondrial genomes; however, all BER proteins are encoded by nuclear genes, and expression of mitochondrial BER (mtBER)protein variants involves alternative transcription initiation sites and alternative splice sites. Some mtBER proteins contain a canonical mitochondrial leader sequence [11, 12], while others localize to the mitochondria by other mechanisms [13].
The expression and enzymatic activity of BER proteins are regulated in a tissue-specific [14, 15], and cell cycle-specific manner[6, 16, 17]. Post-translation modifications [18], and protein-protein interactions [5, 19, 20] also modulate many BER enzyme functions. In the brain, the expression and the activity of BER proteins varies by brain region[21], but the significance of these differences is not well understood.
BER deficiency and neurodegenerative disease
In the mouse, null alleles of genes encoding several core BER proteins are lethal during embryonic development, indicating a critical early developmental role for BER in the mouse. In humans, rare cases of inherited defects in genes encoding BER proteins cause immune system dysfunction [22, 23], cancer[24], and some polymorphic variants seem to influence susceptibility to some human diseases (reviewed in [25, 26]). In addition, defects in BER proteins or in proteins that modulate BER are implicated in several neurodegenerative disorders [27, 28].
For the purpose of this review, we discuss different human disease syndromes linked to defects in BER and other DNA repair pathways. One group include syndromes caused by defects in gene products or genes encoding Polynucleotide kinase phosphatase(PNKP), Ataxia with oculomotor apraxia 1 (APTX) or Tyrosyl-DNA Phosphodiesterase 1 (TDP1) (Fig. 1). Other diseases are genetically and clinically complex, often further complicated with potential environmental triggers or predisposing factors: examples include Parkinson’s disease (PD), Huntington’s Disease (HD) and Alzheimer’s Disease (AD).
Polynucleotide kinase phosphatase (PNKP)
Oxidative damage to DNA by high energy reactive oxygen species (ROS)and γ-radiation can generate DNA single-strand breaks with damaged 5′- and 3′-termini including 5′-OH, and 3′-PO4 that block DNA polymerase repair synthesis and DNA ligation. In BER, the bi-functional DNA glycosylases NEIL1 and NEIL2 possess an AP-lyase activity that produces a 3′-PO4 group by a β, δ-elimination reaction which blocks ligation[29]. PNKP is a bi-functional enzyme with 5′-kinase and 3′-phosphatase activities and is the primary DNA repair protein for removing 5′-OH and 3′-PO4 blocking groups.
In the nucleus, PNKP forms a complex with NEIL1, NEIL2, Polβ, and Lig III [30, 31]. PNKP has also been detected in mitochondria [32, 33], in close proximity to NEIL2 and Polγ [33]. PNKP is also the major 3′-phosphatase in mitochondrial extracts, and PNKP depleted mitochondrial extract showed reduced mtBER activity [33]. These results indicate that PNKP actively participates in nuclear and mtBER.
Mutations in the phosphatase and kinase domains of PNKP have been identified in patients with microcephaly, early-onset seizures and developmental delay (denoted MCSZ Online Mendalian Inheritance in Man (OMIM) # 613402) [34]. A group of disorders of varying severity may be linked to defects in PNKP, as observed for other DNA repair deficiency diseases in which unique mutations give rise to different clinical presentations [35].
Ataxia with oculomotor apraxia 1 (AOA1)
During DNA ligation, a transient covalent bond is formed between AMP and a lysine in the ligase enzyme active site. The AMP moiety, derived from ATP, is then transferred to the 5′P of the DNA substrate, forming 5′-AMP-DNA. The transient 5′-AMP moiety then undergoes nucleophilic attack by the deoxyribosyl 3′ OH, displacing AMP and facilitating ligation of the two DNA strands[36]. Premature termination of DNA ligation generates 5′-AMP termini. The enzyme that removes 5′-AMP from DNA is Aprataxin (APTX) [37]. APTX also removes AMP from 5′-AMP-dRP residues that arise in DNA after attempts to ligate ends with 5′-dRP groups [38].
Mutations in APTX cause the inherited disease ataxia with oculomotor apraxia 1 (AOA1, OMIM #208920) [39, 40]. AOA1 patients are not susceptible to cancer nordo they develop immunological deficiencies, which is consistent with the findings that APTX deficient cells are not distinctly more sensitive togenotoxins than APTX proficient cells [41–43]. APTX localizes to the nuclear and mitochondrial compartments of human cells[43]. Knocking down APTX in SH-SY5Y cells caused loss of mtDNA integrity, but had no detectable effects on the nuclear genome [43]. This suggests a mitochondrial component in the pathogenesis of AOA1. More recently, it has been reported that APTX may play a role in repair of adenylation adducts at RNA-DNA junctions that are likely to be abundant in normal DNA metabolism [44]. Further studies will clarify the role of APTX in the nuclear and mtBER and its relation to AOA1 pathology.
Tyrosyl-DNA phosphodiesterase 1 (TDP1)
Mutations in TDP1 give rise to Spinocerebellarataxia with axonal neuropathy (SCAN1, OMIM # 607250, [45]), characterized by cerebellar atrophy and peripheral neuropathy. TDP1 is important for the repair of 3′phosphoglycolate, 3′abasic sites, topoisomerase I (TOPO1)-linked DNA adducts and others [46]. During DNA replication or transcription, torsional stress is introduced into DNA and TOPO1 makes a transient SSB in the DNA to relieve this stress. During its catalytic cycle, TOPO1 makes a covalent bond between an active site tyrosine residue and the 3′ side of the SSB. If the enzyme is unable to re-ligate the SSB, TOPO1 becomes stuck and the DNA-protein adduct is then subjected to repair by TDP1[45, 47]. TDP1 is localized to both the nucleus and mitochondria [48]. Tdp1−/− mice show age-dependent and progressive cerebellar atrophy[49]. A knockout of the TDP1 ortholog in Drosophila reduced lifespan of female flies, which was rescued by expression of TDP1 in neurons[50].
A high level of DNA damage is reported as a cellular phenotype of the disorders MCSZ, AOA1 and SCAN1. MCSZ and AOA1 patients are not susceptible to cancer or immune system dysfunction, MCSZ seems to be a developmental disorder, whereas AOA1 and SCAN1 are progressive neurodegenerative diseases. It is important to note that PNKP, APTX and TDP1 are expressed in mitochondria and features in these diseases may reflect mitochondrial dysfunction. Generation of exclusively nuclear or mitochondrial targeted enzymes should help to clarify this issue.
Poly(ADP)ribose polymerase (PARP1)
Poly(ADP)ribose polymerase 1 (PARP1) utilizes NAD+ to ADP-ribosylate itself and other protein substrates [51]. While it is not a core BER protein, PARP1 is activated by DNA damage and activated PARP1 plays a role in regulating the DNA damage response (DDR). PARP1 substrate proteins include Polβ, XRCC1 and Lig III[52]. There is much effort towards understanding the role of PARP1 in DNA damage and repair and mitochondrial bioenergetics(for recent reviews see [51, 53]). NAD+ is an important cofactor for many enzymes including PARP1 (and PARP2) in response to DNA damage. Under mild stress, PARP1 recruits proteins to damaged sites and facilitates repair. However, if the damage is too great, persistent PARP1 activation leads to overutilization of NAD+ that can substantially alter the cellular NAD+/NADH ratio and the size of the ATP pool, leading to cell death (reviewed in[54]). Conversely, preventing NAD+ depletion protects neurons against excitotoxicity, death by bioenergetics stress, and ischemia [55, 56]. Thus, one might expect that cells derived from DNA repair deficient patients would have a low level of NAD+, compromised cellular bioenergetics and higher cell death. Increased cell death and PARP1 activation has been seen in Ogg1−/− cells exposed to oxidative stress [57], and in animal models lacking both Ogg1 and Mth1 (an 8-oxodGTPase) following treatment with the mitochondrial complex II inhibitor 3-nitropropionic acid [58]. In addition, in cells from individuals with DNA repair deficiency, chronic PARP1 activation promotes energetic and mitochondrial dysregulation and contributes to cellular dysfunction [59, 60].
Cockayne syndrome group B (CSB) protein
The CSB protein is a SWI/SNF-like DNA-dependent ATPase, a class of proteins that modify the structure of DNA surrounding sites of damage to facilitate repair. CSB is implicated in transcription coupled nucleotide excision repair (TCR), a specialized sub-pathway of nucleotide excision repair (NER) and in BER [61, 62]. DNA damage can inhibit the movement of RNA polymerase II (RNA Pol-II) along DNA. CSB binds to stalled RNA Pol-II and recruits DNA repair proteins. CSB-deficient cells show a widespread transcription defect, indicating an important role for CSB in nuclear transcription [63], and likely also in mitochondrial transcription [64]. Mutations in CSB associate with Cockayne syndrome (CS), a rare inherited disease (OMIM #609413). CS patients are not cancer prone but exhibit growth and developmental defects, accelerating aging, and progressive neurodegeneration [65].
NER removes helix-distorting DNA damage, such as thymine dimers formed by UV-light, hence, the UV-light sensitivity of CSB deficient cells [66]. However, it is unlikely that UV-light penetrates the brain and generates the type of lesions that activate PARP1 in CSB patients[59, 60], suggesting other sources of DNA damage may be involved. CSB deficient cells are sensitive to oxidative stress and accumulate oxidative base lesions [67, 68]. CSB stimulates the activity of DNA glycosylases NEIL1 and NEIL2 [69, 70], APE-1 [71], and colocalizes with NEIL2 at sites of stalled transcription [69]. Furthermore, NEIL2 associates with RNA II, and NEIL2-depleted cells accumulate more damage in active genes [72]. All of this together suggests a potential role for CSB in BER, and that the observed persistent PAPR1 activation in CSB brain may be, at least partly, related to BER deficiency. CSB directly interacts with PARP1[73] and we have recently found that CSB can displace PARP1 from DNA and thus may inhibit excessive PARylation [59].
Ataxia telangiectasia mutated (ATM)
Ataxia telangiectasia mutated (ATM) is a serine/threonine protein kinase activated by DNA double-strand breaks (DSBs). ATM is a master regulator of the response to DNA damage. ATM deficient cells are sensitive to and ATM is activated by oxidative stress [74]. Defects in ATM cause ataxia telangiectasia (A-T), an autosomal recessive disease characterized by cancer predisposition, insulin resistance, and neurodegeneration (OMIM #208900).
ATM protein has been detected in mitochondria, where it appears to regulate both mtDNA copy number [75] and the process of mitophagy[75–77]. ATM deficiency downregulates Lig III and lowers BER capacity in the mitochondria [78]. Lig III is the only DNA ligase in mitochondria and is the rate-limiting enzyme in mtBER[79]. These observations suggest a mitochondrial defect in A-T cells and additional studies of mitochondrial function in A-T is warranted.
We have recently shown that A-T has a mitochondrial phenotype as supported by both bioinformatics and in vitro studies. Using a mitochondrial clinical data established by us (www.mitodb.com), analysis of A-T suggests a mitochondrial disease phenotype. We investigated mitochondrial parameters using in ATM cell lines and found significant mitochondrial defects as evidenced by increased intracellular ROS, mitochondrial ROS, and increased mitochondrial membrane potential. ATM also display accumulation of damaged mitochondrial probably due to defective mitophagy [59].
Alzheimer’s disease, Parkinson’s disease and Huntington’s disease are among the most common neurodegenerative disorders worldwide. Although they each have their specific genetic and physiological mechanisms of pathology, oxidative stress, oxidative damage to macromolecules, mitochondrial dysfunction and nuclear and mtDNA damage constitute the central features of these diseases.
Alzheimer’s disease
Alzheimer’s disease (AD) is a neurodegenerative disorder characterized by progressive cognitive and memory decline (OMIM #104300). Sporadic AD with late-on set is more common, while the less prevalent form of AD is early-onset autosomal dominant, linked to mutations in genes encoding amyloid precursor protein (APP), PSEN1 (PS1), and PSEN2 (PS2) [80].
Multiple studies show that oxidative modifications of lipids, proteins, and nucleic acids are higher in brains of AD patients than in normal brain, often accompanied with impaired neuroprotective stress responses[81, 82]. Oxidative stress seems to be a systemic feature of AD and an oxidative stress phenotypehas been seen in AD fibroblasts [83]. High levels of oxidative DNA base lesions have been reported in preclinical [84] and late stage AD brains [85]. Increased amounts of oxidative DNA damage are a manifestation of an imbalance between ROS production (e.g. from dysfunctional mitochondria), and antioxidant and BER capacity of the cell. Lower nuclear and mitochondrial BER activities have been observed in AD brain samples [86–88]. Mitochondrial dysfunction and increased mtDNA damage have been reported frequently in AD brains [89–91]. 5-hydroxyuracil (5OHU) incision and DNA ligation activities were significantly reduced in mitochondrial lysates from AD brains [87]. In the same samples, the concentration of NEIL1 DNA glycosylase was much lower than in the control samples [87]. Together, these results support the idea that increased mtDNA damage together with a reduction of mtBER play a role in AD pathology [82, 89].
Interestingly, the expression and activity of Polβ were considerably reduced in brains of mild cognitive impaired (MCI) and AD patients, and showed an inverse correlation with the severity of the disease [86]. A reduction of Polβ is also seen in Down syndrome (DS)[92]. DS brains exhibit high levels of ROS and have significant oxidative damage [93], and DS patients develop AD at very high rates. Thus, increased oxidative DNA damage concomitant with a reduction in BER capacity seem to be a common feature of DS and AD. For these reasons, a novel mouse model combining the 3xTg AD mouse with a Polβ+/− heterozygote was created to specifically test if loss of Polβ alters the timing or presentation of symptoms in this AD mouse[94]. Interestingly, the 3xTg AD/Polβ+/− mouse showed greater cell death, neuronal loss and impaired mitochondrial bioenergetics, thus demonstrating that reduced BER capacity can substantially exacerbate disease progression. Polβ deficient AD mice and human AD patients displayed mitochondrial dysfunction and altered gene expression, while similar patterns were not detected in AD mice. Polβ depletion also appeared to cause defects in memory and neurogenesis [94].
Parkinson’s disease
Parkinson’s disease (PD) is a progressive neurodegenerative disease characterized by motor and non-motor symptoms and the loss of dopaminergic neurons in substantianigrain the midbrain (OMIM #168600). Oxidative damage to macromolecules is an intrinsic property of PD. ROS generated as a by-product of dopamine metabolism, is considered a significant contributor todopaminergic neuronal loss in the PD brain [95].
A mitochondrial deficiency in PD pathogenesis, especially in the substantianigra, is extensively documented[96–98]. Environmental toxins like rotenone that inhibit mitochondrial complex I, have long been known to cause pathological features of PD in humans and in animals. Exposure of mice to rotenone impaired mitochondrial complex I function and damaged mtDNA before the PD symptoms were manifested [99]. In a separate study, a significantly increased level of mtDNA mutations was identified in early PD cases [100]. These results suggest that loss of mtDNA integrity is among the early events in the onset of PD.
Huntington’s disease
Huntington’s disease (HD) is an autosomal dominant hereditary neurodegenerative disease characterized by progressive decline in motor and cognitive functions (OMIM #143100). HD is one of over 40 neurodegenerative disorders caused by the expansion of trinucleotide repeat (TNR) sequences in DNA. Huntingtin gene (HTT) codes for the huntingtin protein (Htt). The function of Htt is not completely understood, but it is essential for normal embryonic development and neurogenesis, and its complete knockout in mice is lethal[101]. The HTT gene has an unstable CAG repeat in its first exon and CAG expansion is especially significant in the striatum, the brain region that degenerates in HD. An expansion of 40 or more of the CAG repeats causes HD and the clinical severity of the disease directly correlates with the length of the expansion [102]. BER of oxidized DNA bases has been implicated in CAG expansion in HD. In support of this, loss of Ogg1 DNA glycosylase suppressed CAG expansion in HD mice [103].
A key role for mitochondrial dysfunction in HD neuropathology is well documented [104–107]. Moreover, HD brains show markers of enhanced oxidative stress including oxidative damage to nuclear and mitochondrial genomes [108–110], and targeting antioxidant to mitochondria reduced mtDNA damage and suppressed motor decline in a mouse model of HD [111]. Given the observed increased oxidative stress and the abnormalities in mitochondrial bioenergetics in HD, more studies on the state of mtBER in HD are warranted.
BER in ischemia and stroke
Ischemic-reperfusion (I/R) occurs when the blood supply to an organ is disrupted and then restored, as in stroke. Reperfusion can cause a sudden increase in free radicals causing tissue injury and cell death, for instance, following heart attack and stroke. Studies in mice suggest that BER is important for recovery from ischemic-reperfusion injury; however, some studies suggest that active BER may promote brain damage during ischemia[112].
Increased expression of mtBER and nuclear BER proteins in response to focal ischemia have been reported[113, 114], with the level of tissue and oxidative DNA base damage inversely correlating with BER capacity [114, 115]. In addition, the size of the brain infarct was reported to be larger in Ung, Ogg1, Neil1 and Neil3 null mice than in wild type mice after experimentally-induced stroke[116–119]. Conversely, depletion of alkyladenine DNA glycosylase (Aag1/Mpg), has a protective affect after induced stroke/ischemia [112]. Loss of Aag1 caused reduced PARP1 activation and less NAD+ depletion. Moreover, at least after liver ischemic injury, the detrimental effects of Aag1 could be offset by Parp1 knockout, further supporting the idea that PARP1 activation was the key. Clearly more research is necessary to understand whether and how BER capacity and activity influences stroke outcome; and also whether PARP1 inhibitors and exogenous NAD+ have potential therapeutic applications for stroke patients.
Vascular dementia is a type of cognitive impairment thought to be caused by vascular damage from small strokes [120], that may contribute to pathology in AD patients [121]. Examination of BER null mice brains for small strokes and microvascular pathology may clarify the relationship between oxidative DNA damage and DNA repair in onset and progression of neurodegenerative disorders and AD.
PERSPECTIVES
Extensive experimental evidence suggests that oxidative stress, mitochondrial dysfunction, and persistent DNA damage play roles in onset and progression of human neurodegenerative disease. Although we do not yet understand in detail how tissue- and organelle-specific BER affects the risk of neurodegenerative disease, it is a subject of much research. Several promising but preliminary studies in this research area are mentioned briefly below; we also list significant roadblocks and bottlenecks inhibiting research in the field, and new research ideas relevant to BER related neurological disease.
Behavioral studies in Neil1 deficient mice revealed specific impairment of some memory functions, and defects in olfaction[119, 122]. As discussed earlier, the level of NEIL1 appears to be significantly lower in AD brain than in normal brain[122]. It is interesting to note that a poor sense of smell is considered an early sign of risk for cognitive decline and AD [123]. These results suggest a possible role for NEIL1 in brain homeostasis. It will be interesting to perform similar studies on mice lacking other DNA glycosylases as well as to examine other sensorineural functions in Neil-1 deficient mice. Notably, CSB deficient mice are defective in sensorineural hearing[60], which is considered to be a feature of normal aging in humans.
Post-mortem human brains are a very valuable resource for studying human brain neurodegeneration. However, many questions cannot be addressed using such tissue, nor can they be addressed by descriptive retrospective studies of diseased and normal brains. For example, it has been proposed that BER dynamics may modulate susceptibility to neurodegenerative disease. Such dynamics can only be addressed currently in animal- or cell-based model systems. If one proposes to study quantify or characterize specific BER enzyme functions in post-mortem brain, the amount of tissue available can be prohibitively small. For example, it is a big challenge to get enough and highly purified material to study mtBER activity in postmortem brain tissue [87]. Alternative approaches could include future use of induced pluripotent stem cells (iPSCs), where such cells could even carry putative AD susceptibility alleles. This technique has been successfully used to study sporadic and familial AD [124]. Future use of this technology could make it possible to test the hypothesis that BER capacity correlates with risk for neurological dysfunction.
Expression of BER proteins varies by brain region, brain cell type, and fluctuates with age[125]. BER imbalance can generate DNA repair intermediates that are mutagenic and cytotoxic [126], and contribute to oxidative stress-induced neuronal loss [127]. Caenorhabditiselegans and Drosophila melanogaster might provide model systems to explore whether and how BER imbalance influences neurological function and animal behavior.
It has been proposed that somatic mutational load may correlate with susceptibility to neurodegenerative disease, possibly because somatic mutations can give rise to mutated proteins [128], activate the cellular DNA damage response and lead to apoptotic cell death. Because BER capacity appears to be lower in mitochondria than in the nucleus, lesions in mtDNA could persist longer and have more deleterious consequences than lesions in the nuclear DNA. Thus, quantitative analysis of steady state level of mtDNA damage is a critical area of future study. New technologies or established technology with improved sensitivity and specificity (i.e., liquid chromatography-tandem mass spectrometry) may help advance such studies [129].
Recent studies suggest that CSB modulates mitophagy (i.e., autophagic removal of damaged mitochondria) [130], and it has been suggested that persistent PARP1 activity, NAD+ depletion and mitochondrial dysfunction may contribute to pathology in XPA, ATM and CS [59]. We proposed that this reflects nuclear mitochondrial signaling mediated by PARylation. This idea was tested by treating C. elegans and mice with exogenous NAD+ or PARP inhibitors, which seemed to restore mitochondrial function and mitigate other disease symptoms in these animal model systems[60]. As mentioned earlier in this review, additional study of exogenous NAD+ and PARP inhibitors as therapeutic agents for neurodegenerative diseases is warranted.
Alzheimer’s disease is a major social and medical challenge. Our understanding of disease progression is limited and no effective treatments are yet available [131]. Studies discussed in this review[81–85, 91]support the notion that oxidative stress as well as deficient repair of oxidative damage contribute to the pathology of AD. Useful insight might emerge from experiments in which different DNA repair deficient genotypes were crossed into representative AD mouse models, followed by phenotyping of double heterozygotes or double heterozygous mice. Phenotyping should investigate physiological and behavioral characteristics of the novel mice strains as well as DNA repair capacity in brain and other tissues. Tau phosphorylation and pathology is associated with oxidative stress and possibly with mitochondrial function (reviewed in [132]). It would be of interest to explore whether there is a connection with DNA repair processes in mitochondria and Tau pathology. While a BER defect may not be the causative issue in AD, deficient BER may promote the progression of the disease [94], and thus be a significant risk factor in AD.
In our opinion, BER is a worthy target for interventions (Table I) but we also recognize that BER is a highly regulated and coordinated process. Thus, it is unlikely that overexpression of any one BER enzyme in an unregulated manner would be beneficial, even in BER deficient cells [133]. Furthermore, there is no general agreement as to which step of BER is rate-limiting; and the rate-limiting step of BER might be different in different cellular microenvironments, tissues or organelles. Another approach worthy of consideration might be to indirectly stimulate BER by modulating rates of protein ubiquitination [134, 135]or by expressing regulatory microRNA species in a controlled manner[136]. Additionally, the identification and manipulation of microRNAs that regulate multiple proteins in the pathway could represent another means of upregulating the entire pathway.
Table I.
Points of Intervention
| Up-regulation | Down-regulation |
|---|---|
| DNA glycosylases | PARP1 |
| DNA polymerase β | |
| DNA ligase III | |
| NAD+ levels | NAD+ consumption |
Acknowledgments
We thank Dr. Mimi Sander, Page One Editorial Services for editing. We thank Thomas Wynn for assistance with graphics. We to thank Drs. Evandro Fang and Beverly Baptiste for critical comments on the article. This research was supported in part by the Intramural Research Program of the NIH, National Institute on Aging.
ABBREVIATIONS
- AD
Alzheimer’s Disease
- AMP
Adenosine monophosphate
- AOA1
Ataxia with oculomotor apraxia 1
- APE-1
AP endonuclease 1
- APP
Amyloid precursor protein
- APTX
Aprataxin
- ATM
Ataxia telangiectasia mutated
- BER
DNA base excision repair
- CS
Cockayne syndrome
- CSB
Cockayne syndrome group B protein
- DDR
DNA damage response
- 5′-dRP
5′-Deoxyribose phosphate
- DSB
DNA double-strand breaks
- HD
Huntington’s Disease
- Lig III
DNA ligase III
- MCI
Mild cognitive impaired
- MCSZ
Microcephaly, early-onset seizures and developmental delay
- mtBER
Mitochondrial base excision repair
- mtDNA
Mitochondrial DNA
- NAD
Nicotinamide adenine dinucleotide
- NEIL1
Endonuclease VIII-like
- NER
Nucleotide excision repair
- OGG1
8-oxoguanine DNA glycosylase
- PARP1
Poly(ADP)ribose polymerase 1
- PARylation
Poly(ADP-ribosyl)ation
- PD
Parkinson’s Disease
- PNKP
Polynucleotide kinase phosphatase
- Polβ
DNA polymerase β
- PSEN
Presenilin
- RNA Pol
RNA polymerase
- ROS
Reactive oxygen species
- SCAN1
Spinocerebellarataxia with axonal neuropathy
- SSB
DNA single-strand break
- TDP1
Tyrosyl-DNA phosphodiesterase 1
- TOPO1
Topoisomerase I
- UNG
Uracil-DNA glycosylase
- XRCC1
X-ray repair cross-complementing protein 1
Footnotes
Publisher's Disclaimer: This is a PDF file of an unedited manuscript that has been accepted for publication. As a service to our customers we are providing this early version of the manuscript. The manuscript will undergo copyediting, typesetting, and review of the resulting proof before it is published in its final citable form. Please note that during the production process errors may be discovered which could affect the content, and all legal disclaimers that apply to the journal pertain.
References
- 1.Salminen LE, Paul RH. Oxidative stress and genetic markers of suboptimal antioxidant defense in the aging brain: a theoretical review. Reviews in the neurosciences. 2014;25:805–819. doi: 10.1515/revneuro-2014-0046. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 2.Lauritzen KH, Moldestad O, Eide L, et al. Mitochondrial DNA toxicity in forebrain neurons causes apoptosis, neurodegeneration, and impaired behavior. Molecular and cellular biology. 2010;30:1357–1367. doi: 10.1128/MCB.01149-09. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 3.Yang JL, Tadokoro T, Keijzers G, et al. Neurons efficiently repair glutamate-induced oxidative DNA damage by a process involving CREB-mediated up-regulation of apurinic endonuclease 1. The Journal of biological chemistry. 2010;285:28191–28199. doi: 10.1074/jbc.M109.082883. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 4.Suberbielle E, Sanchez PE, Kravitz AV, et al. Physiologic brain activity causes DNA double-strand breaks in neurons, with exacerbation by amyloid-beta. Nature neuroscience. 2013;16:613–621. doi: 10.1038/nn.3356. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 5.Fan J, Wilson DM., 3rd Protein-protein interactions and posttranslational modifications in mammalian base excision repair. Free radical biology & medicine. 2005;38:1121–1138. doi: 10.1016/j.freeradbiomed.2005.01.012. [DOI] [PubMed] [Google Scholar]
- 6.Sykora P, Yang JL, Ferrarelli LK, et al. Modulation of DNA base excision repair during neuronal differentiation. Neurobiology of aging. 2013;34:1717–1727. doi: 10.1016/j.neurobiolaging.2012.12.016. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 7.Sjolund AB, Senejani AG, Sweasy JB. MBD4 and TDG: multifaceted DNA glycosylases with ever expanding biological roles. Mutation research. 2013;743–744:12–25. doi: 10.1016/j.mrfmmm.2012.11.001. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 8.Rada C, Williams GT, Nilsen H, et al. Immunoglobulin isotype switching is inhibited and somatic hypermutation perturbed in UNG-deficient mice. Current biology : CB. 2002;12:1748–1755. doi: 10.1016/s0960-9822(02)01215-0. [DOI] [PubMed] [Google Scholar]
- 9.Jobert L, Skjeldam HK, Dalhus B, et al. The human base excision repair enzyme SMUG1 directly interacts with DKC1 and contributes to RNA quality control. Molecular cell. 2013;49:339–345. doi: 10.1016/j.molcel.2012.11.010. [DOI] [PubMed] [Google Scholar]
- 10.Krokan HE, Saetrom P, Aas PA, et al. Error-free versus mutagenic processing of genomic uracil--relevance to cancer. DNA repair. 2014;19:38–47. doi: 10.1016/j.dnarep.2014.03.028. [DOI] [PubMed] [Google Scholar]
- 11.Nilsen H, Otterlei M, Haug T, et al. Nuclear and mitochondrial uracil-DNA glycosylases are generated by alternative splicing and transcription from different positions in the UNG gene. Nucleic acids research. 1997;25:750–755. doi: 10.1093/nar/25.4.750. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 12.Lakshmipathy U, Campbell C. The human DNA ligase III gene encodes nuclear and mitochondrial proteins. Molecular and cellular biology. 1999;19:3869–3876. doi: 10.1128/mcb.19.5.3869. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 13.Li M, Zhong Z, Zhu J, et al. Identification and characterization of mitochondrial targeting sequence of human apurinic/apyrimidinic endonuclease 1. The Journal of biological chemistry. 2010;285:14871–14881. doi: 10.1074/jbc.M109.069591. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 14.Karahalil B, Hogue BA, de Souza-Pinto NC, Bohr VA. Base excision repair capacity in mitochondria and nuclei: tissue-specific variations. FASEB journal : official publication of the Federation of American Societies for Experimental Biology. 2002;16:1895–1902. doi: 10.1096/fj.02-0463com. [DOI] [PubMed] [Google Scholar]
- 15.Haug T, Skorpen F, Aas PA, et al. Regulation of expression of nuclear and mitochondrial forms of human uracil-DNA glycosylase. Nucleic acids research. 1998;26:1449–1457. doi: 10.1093/nar/26.6.1449. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 16.Hardeland U, Kunz C, Focke F, et al. Cell cycle regulation as a mechanism for functional separation of the apparently redundant uracil DNA glycosylases TDG and UNG2. Nucleic acids research. 2007;35:3859–3867. doi: 10.1093/nar/gkm337. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 17.Narciso L, Fortini P, Pajalunga D, et al. Terminally differentiated muscle cells are defective in base excision DNA repair and hypersensitive to oxygen injury. Proceedings of the National Academy of Sciences of the United States of America. 2007;104:17010–17015. doi: 10.1073/pnas.0701743104. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 18.Hagen L, Kavli B, Sousa MM, et al. Cell cycle-specific UNG2 phosphorylations regulate protein turnover, activity and association with RPA. The EMBO journal. 2008;27:51–61. doi: 10.1038/sj.emboj.7601958. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 19.Hanssen-Bauer A, Solvang-Garten K, Sundheim O, et al. XRCC1 coordinates disparate responses and multiprotein repair complexes depending on the nature and context of the DNA damage. Environmental and molecular mutagenesis. 2011;52:623–635. doi: 10.1002/em.20663. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 20.Akbari M, Solvang-Garten K, Hanssen-Bauer A, et al. Direct interaction between XRCC1 and UNG2 facilitates rapid repair of uracil in DNA by XRCC1 complexes. DNA repair. 2010;9:785–795. doi: 10.1016/j.dnarep.2010.04.002. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 21.Rolseth V, Runden-Pran E, Luna L, et al. Widespread distribution of DNA glycosylases removing oxidative DNA lesions in human and rodent brains. DNA repair. 2008;7:1578–1588. doi: 10.1016/j.dnarep.2008.06.007. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 22.Imai K, Slupphaug G, Lee WI, et al. Human uracil-DNA glycosylase deficiency associated with profoundly impaired immunoglobulin class-switch recombination. Nature immunology. 2003;4:1023–1028. doi: 10.1038/ni974. [DOI] [PubMed] [Google Scholar]
- 23.Barnes DE, Tomkinson AE, Lehmann AR, et al. Mutations in the DNA ligase I gene of an individual with immunodeficiencies and cellular hypersensitivity to DNA-damaging agents. Cell. 1992;69:495–503. doi: 10.1016/0092-8674(92)90450-q. [DOI] [PubMed] [Google Scholar]
- 24.Chow E, Thirlwell C, Macrae F, Lipton L. Colorectal cancer and inherited mutations in base-excision repair. The Lancet. Oncology. 2004;5:600–606. doi: 10.1016/S1470-2045(04)01595-5. [DOI] [PubMed] [Google Scholar]
- 25.Wilson DM, 3rd, Kim D, Berquist BR, Sigurdson AJ. Variation in base excision repair capacity. Mutation research. 2011;711:100–112. doi: 10.1016/j.mrfmmm.2010.12.004. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 26.Karahalil B, Bohr VA, Wilson DM., 3rd Impact of DNA polymorphisms in key DNA base excision repair proteins on cancer risk. Human & experimental toxicology. 2012;31:981–1005. doi: 10.1177/0960327112444476. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 27.Sykora P, Wilson DM, 3rd, Bohr VA. Base excision repair in the mammalian brain: implication for age related neurodegeneration. Mechanisms of ageing and development. 2013;134:440–448. doi: 10.1016/j.mad.2013.04.005. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 28.Wallace SS. Base excision repair: a critical player in many games. DNA repair. 2014;19:14–26. doi: 10.1016/j.dnarep.2014.03.030. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 29.Wallace SS. DNA glycosylases search for and remove oxidized DNA bases. Environmental and molecular mutagenesis. 2013;54:691–704. doi: 10.1002/em.21820. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 30.Das A, Wiederhold L, Leppard JB, et al. NEIL2-initiated, APE-independent repair of oxidized bases in DNA: Evidence for a repair complex in human cells. DNA repair. 2006;5:1439–1448. doi: 10.1016/j.dnarep.2006.07.003. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 31.Wiederhold L, Leppard JB, Kedar P, et al. AP endonuclease-independent DNA base excision repair in human cells. Molecular cell. 2004;15:209–220. doi: 10.1016/j.molcel.2004.06.003. [DOI] [PubMed] [Google Scholar]
- 32.Tahbaz N, Subedi S, Weinfeld M. Role of polynucleotide kinase/phosphatase in mitochondrial DNA repair. Nucleic acids research. 2012;40:3484–3495. doi: 10.1093/nar/gkr1245. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 33.Mandal SM, Hegde ML, Chatterjee A, et al. Role of human DNA glycosylase Nei-like 2 (NEIL2) and single strand break repair protein polynucleotide kinase 3′-phosphatase in maintenance of mitochondrial genome. The Journal of biological chemistry. 2012;287:2819–2829. doi: 10.1074/jbc.M111.272179. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 34.Shen J, Gilmore EC, Marshall CA, et al. Mutations in PNKP cause microcephaly, seizures and defects in DNA repair. Nature genetics. 2010;42:245–249. doi: 10.1038/ng.526. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 35.DiGiovanna JJ, Kraemer KH. Shining a light on xeroderma pigmentosum. The Journal of investigative dermatology. 2012;132:785–796. doi: 10.1038/jid.2011.426. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 36.Ellenberger T, Tomkinson AE. Eukaryotic DNA ligases: structural and functional insights. Annual review of biochemistry. 2008;77:313–338. doi: 10.1146/annurev.biochem.77.061306.123941. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 37.Ahel I, Rass U, El-Khamisy SF, et al. The neurodegenerative disease protein aprataxin resolves abortive DNA ligation intermediates. Nature. 2006;443:713–716. doi: 10.1038/nature05164. [DOI] [PubMed] [Google Scholar]
- 38.Caglayan M, Batra VK, Sassa A, et al. Role of polymerase beta in complementing aprataxin deficiency during abasic-site base excision repair. Nature structural & molecular biology. 2014;21:497–499. doi: 10.1038/nsmb.2818. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 39.Date H, Onodera O, Tanaka H, et al. Early-onset ataxia with ocular motor apraxia and hypoalbuminemia is caused by mutations in a new HIT superfamily gene. Nature genetics. 2001;29:184–188. doi: 10.1038/ng1001-184. [DOI] [PubMed] [Google Scholar]
- 40.Moreira MC, Barbot C, Tachi N, et al. The gene mutated in ataxia-ocular apraxia 1 encodes the new HIT/Zn-finger protein aprataxin. Nature genetics. 2001;29:189–193. doi: 10.1038/ng1001-189. [DOI] [PubMed] [Google Scholar]
- 41.Clements PM, Breslin C, Deeks ED, et al. The ataxia-oculomotor apraxia 1 gene product has a role distinct from ATM and interacts with the DNA strand break repair proteins XRCC1 and XRCC4. DNA repair. 2004;3:1493–1502. doi: 10.1016/j.dnarep.2004.06.017. [DOI] [PubMed] [Google Scholar]
- 42.El-Khamisy SF, Katyal S, Patel P, et al. Synergistic decrease of DNA single-strand break repair rates in mouse neural cells lacking both Tdp1 and aprataxin. DNA repair. 2009;8:760–766. doi: 10.1016/j.dnarep.2009.02.002. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 43.Sykora P, Croteau DL, Bohr VA, Wilson DM., 3rd Aprataxin localizes to mitochondria and preserves mitochondrial function. Proceedings of the National Academy of Sciences of the United States of America. 2011;108:7437–7442. doi: 10.1073/pnas.1100084108. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 44.Tumbale P, Williams JS, Schellenberg MJ, et al. Aprataxin resolves adenylated RNA-DNA junctions to maintain genome integrity. Nature. 2014;506:111–115. doi: 10.1038/nature12824. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 45.Takashima H, Boerkoel CF, John J, et al. Mutation of TDP1, encoding a topoisomerase I-dependent DNA damage repair enzyme, in spinocerebellar ataxia with axonal neuropathy. Nature genetics. 2002;32:267–272. doi: 10.1038/ng987. [DOI] [PubMed] [Google Scholar]
- 46.Pommier Y, Huang SY, Gao R, et al. Tyrosyl-DNA-phosphodiesterases (TDP1 and TDP2) DNA repair. 2014;19:114–129. doi: 10.1016/j.dnarep.2014.03.020. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 47.El-Khamisy SF, Saifi GM, Weinfeld M, et al. Defective DNA single-strand break repair in spinocerebellar ataxia with axonal neuropathy-1. Nature. 2005;434:108–113. doi: 10.1038/nature03314. [DOI] [PubMed] [Google Scholar]
- 48.Das BB, Dexheimer TS, Maddali K, Pommier Y. Role of tyrosyl-DNA phosphodiesterase (TDP1) in mitochondria. Proceedings of the National Academy of Sciences of the United States of America. 2010;107:19790–19795. doi: 10.1073/pnas.1009814107. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 49.Katyal S, el-Khamisy SF, Russell HR, et al. TDP1 facilitates chromosomal single-strand break repair in neurons and is neuroprotective in vivo. The EMBO journal. 2007;26:4720–4731. doi: 10.1038/sj.emboj.7601869. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 50.Guo D, Dexheimer TS, Pommier Y, Nash HA. Neuroprotection and repair of 3′-blocking DNA ends by glaikit (gkt) encoding Drosophila tyrosyl-DNA phosphodiesterase 1 (TDP1) Proceedings of the National Academy of Sciences of the United States of America. 2014;111:15816–15820. doi: 10.1073/pnas.1415011111. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 51.Fouquerel E, Sobol RW. ARTD1 (PARP1) activation and NAD(+) in DNA repair and cell death. DNA repair. 2014;23:27–32. doi: 10.1016/j.dnarep.2014.09.004. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 52.Okano S, Lan L, Caldecott KW, et al. Spatial and temporal cellular responses to single-strand breaks in human cells. Molecular and cellular biology. 2003;23:3974–3981. doi: 10.1128/MCB.23.11.3974-3981.2003. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 53.Guarente L. Linking DNA Damage, NAD(+)/SIRT1, and Aging. Cell metabolism. 2014;20:706–707. doi: 10.1016/j.cmet.2014.10.015. [DOI] [PubMed] [Google Scholar]
- 54.Imai S, Guarente L. NAD+ and sirtuins in aging and disease. Trends in cell biology. 2014;24:464–471. doi: 10.1016/j.tcb.2014.04.002. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 55.Liu D, Pitta M, Mattson MP. Preventing NAD(+) depletion protects neurons against excitotoxicity: bioenergetic effects of mild mitochondrial uncoupling and caloric restriction. Annals of the New York Academy of Sciences. 2008;1147:275–282. doi: 10.1196/annals.1427.028. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 56.Liu D, Gharavi R, Pitta M, et al. Nicotinamide prevents NAD+ depletion and protects neurons against excitotoxicity and cerebral ischemia: NAD+ consumption by SIRT1 may endanger energetically compromised neurons. Neuromolecular medicine. 2009;11:28–42. doi: 10.1007/s12017-009-8058-1. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 57.Oka S, Ohno M, Tsuchimoto D, et al. Two distinct pathways of cell death triggered by oxidative damage to nuclear and mitochondrial DNAs. The EMBO journal. 2008;27:421–432. doi: 10.1038/sj.emboj.7601975. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 58.Sheng Z, Oka S, Tsuchimoto D, et al. 8-Oxoguanine causes neurodegeneration during MUTYH-mediated DNA base excision repair. The Journal of clinical investigation. 2012;122:4344–4361. doi: 10.1172/JCI65053. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 59.Fang EF, Scheibye-Knudsen M, Brace LE, et al. Defective mitophagy in XPA via PARP-1 hyperactivation and NAD(+)/SIRT1 reduction. Cell. 2014;157:882–896. doi: 10.1016/j.cell.2014.03.026. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 60.Scheibye-Knudsen M, Mitchell SJ, Fang EF, et al. A High-Fat Diet and NAD(+) Activate Sirt1 to Rescue Premature Aging in Cockayne Syndrome. Cell metabolism. 2014;20:840–855. doi: 10.1016/j.cmet.2014.10.005. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 61.Kamenisch Y, Berneburg M. Mitochondrial CSA and CSB: protein interactions and protection from ageing associated DNA mutations. Mechanisms of ageing and development. 2013;134:270–274. doi: 10.1016/j.mad.2013.03.005. [DOI] [PubMed] [Google Scholar]
- 62.Aamann MD, Muftuoglu M, Bohr VA, Stevnsner T. Multiple interaction partners for Cockayne syndrome proteins: implications for genome and transcriptome maintenance. Mechanisms of ageing and development. 2013;134:212–224. doi: 10.1016/j.mad.2013.03.009. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 63.Kristensen U, Epanchintsev A, Rauschendorf MA, et al. Regulatory interplay of Cockayne syndrome B ATPase and stress-response gene ATF3 following genotoxic stress. Proceedings of the National Academy of Sciences of the United States of America. 2013;110:E2261–2270. doi: 10.1073/pnas.1220071110. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 64.Berquist BR, Canugovi C, Sykora P, et al. Human Cockayne syndrome B protein reciprocally communicates with mitochondrial proteins and promotes transcriptional elongation. Nucleic acids research. 2012;40:8392–8405. doi: 10.1093/nar/gks565. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 65.Frontini M, Proietti-De-Santis L. Interaction between the Cockayne syndrome B and p53 proteins: implications for aging. Aging. 2012;4:89–97. doi: 10.18632/aging.100439. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 66.Marteijn JA, Lans H, Vermeulen W, Hoeijmakers JH. Understanding nucleotide excision repair and its roles in cancer and ageing. Nature reviews. Molecular cell biology. 2014;15:465–481. doi: 10.1038/nrm3822. [DOI] [PubMed] [Google Scholar]
- 67.Trapp C, Reite K, Klungland A, Epe B. Deficiency of the Cockayne syndrome B (CSB) gene aggravates the genomic instability caused by endogenous oxidative DNA base damage in mice. Oncogene. 2007;26:4044–4048. doi: 10.1038/sj.onc.1210167. [DOI] [PubMed] [Google Scholar]
- 68.Tuo J, Jaruga P, Rodriguez H, et al. Primary fibroblasts of Cockayne syndrome patients are defective in cellular repair of 8-hydroxyguanine and 8-hydroxyadenine resulting from oxidative stress. FASEB journal : official publication of the Federation of American Societies for Experimental Biology. 2003;17:668–674. doi: 10.1096/fj.02-0851com. [DOI] [PubMed] [Google Scholar]
- 69.Aamann MD, Hvitby C, Popuri V, et al. Cockayne Syndrome group B protein stimulates NEIL2 DNA glycosylase activity. Mechanisms of ageing and development. 2014;135:1–14. doi: 10.1016/j.mad.2013.12.008. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 70.Muftuoglu M, de Souza-Pinto NC, Dogan A, et al. Cockayne syndrome group B protein stimulates repair of formamidopyrimidines by NEIL1 DNA glycosylase. The Journal of biological chemistry. 2009;284:9270–9279. doi: 10.1074/jbc.M807006200. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 71.Wong HK, Muftuoglu M, Beck G, et al. Cockayne syndrome B protein stimulates apurinic endonuclease 1 activity and protects against agents that introduce base excision repair intermediates. Nucleic acids research. 2007;35:4103–4113. doi: 10.1093/nar/gkm404. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 72.Banerjee D, Mandal SM, Das A, et al. Preferential repair of oxidized base damage in the transcribed genes of mammalian cells. The Journal of biological chemistry. 2011;286:6006–6016. doi: 10.1074/jbc.M110.198796. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 73.Thorslund T, von Kobbe C, Harrigan JA, et al. Cooperation of the Cockayne syndrome group B protein and poly(ADP-ribose) polymerase 1 in the response to oxidative stress. Molecular and cellular biology. 2005;25:7625–7636. doi: 10.1128/MCB.25.17.7625-7636.2005. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 74.Guo Z, Kozlov S, Lavin MF, et al. ATM activation by oxidative stress. Science. 2010;330:517–521. doi: 10.1126/science.1192912. [DOI] [PubMed] [Google Scholar]
- 75.Eaton JS, Lin ZP, Sartorelli AC, et al. Ataxia-telangiectasia mutated kinase regulates ribonucleotide reductase and mitochondrial homeostasis. The Journal of clinical investigation. 2007;117:2723–2734. doi: 10.1172/JCI31604. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 76.Ambrose M, Goldstine JV, Gatti RA. Intrinsic mitochondrial dysfunction in ATM-deficient lymphoblastoid cells. Human molecular genetics. 2007;16:2154–2164. doi: 10.1093/hmg/ddm166. [DOI] [PubMed] [Google Scholar]
- 77.Valentin-Vega YA, Maclean KH, Tait-Mulder J, et al. Mitochondrial dysfunction in ataxia-telangiectasia. Blood. 2012;119:1490–1500. doi: 10.1182/blood-2011-08-373639. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 78.Sharma NK, Lebedeva M, Thomas T, et al. Intrinsic mitochondrial DNA repair defects in Ataxia Telangiectasia. DNA repair. 2014;13:22–31. doi: 10.1016/j.dnarep.2013.11.002. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 79.Akbari M, Keijzers G, Maynard S, et al. Overexpression of DNA ligase III in mitochondria protects cells against oxidative stress and improves mitochondrial DNA base excision repair. DNA repair. 2014;16:44–53. doi: 10.1016/j.dnarep.2014.01.015. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 80.Guerreiro RJ, Gustafson DR, Hardy J. The genetic architecture of Alzheimer’s disease: beyond APP, PSENs and APOE. Neurobiology of aging. 2012;33:437–456. doi: 10.1016/j.neurobiolaging.2010.03.025. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 81.Lu T, Aron L, Zullo J, et al. REST and stress resistance in ageing and Alzheimer’s disease. Nature. 2014;507:448–454. doi: 10.1038/nature13163. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 82.Wang X, Wang W, Li L, et al. Oxidative stress and mitochondrial dysfunction in Alzheimer’s disease. Biochimica et biophysica acta. 2014;1842:1240–1247. doi: 10.1016/j.bbadis.2013.10.015. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 83.Ramamoorthy M, Sykora P, Scheibye-Knudsen M, et al. Sporadic Alzheimer disease fibroblasts display an oxidative stress phenotype. Free radical biology & medicine. 2012;53:1371–1380. doi: 10.1016/j.freeradbiomed.2012.07.018. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 84.Lovell MA, Soman S, Bradley MA. Oxidatively modified nucleic acids in preclinical Alzheimer’s disease (PCAD) brain. Mechanisms of ageing and development. 2011;132:443–448. doi: 10.1016/j.mad.2011.08.003. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 85.Gabbita SP, Lovell MA, Markesbery WR. Increased nuclear DNA oxidation in the brain in Alzheimer’s disease. Journal of neurochemistry. 1998;71:2034–2040. doi: 10.1046/j.1471-4159.1998.71052034.x. [DOI] [PubMed] [Google Scholar]
- 86.Weissman L, Jo DG, Sorensen MM, et al. Defective DNA base excision repair in brain from individuals with Alzheimer’s disease and amnestic mild cognitive impairment. Nucleic acids research. 2007;35:5545–5555. doi: 10.1093/nar/gkm605. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 87.Canugovi C, Shamanna RA, Croteau DL, Bohr VA. Base excision DNA repair levels in mitochondrial lysates of Alzheimer’s disease. Neurobiology of aging. 2014;35:1293–1300. doi: 10.1016/j.neurobiolaging.2014.01.004. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 88.Lovell MA, Xie C, Markesbery WR. Decreased base excision repair and increased helicase activity in Alzheimer’s disease brain. Brain research. 2000;855:116–123. doi: 10.1016/s0006-8993(99)02335-5. [DOI] [PubMed] [Google Scholar]
- 89.Coskun P, Wyrembak J, Schriner SE, et al. A mitochondrial etiology of Alzheimer and Parkinson disease. Biochimica et biophysica acta. 2012;1820:553–564. doi: 10.1016/j.bbagen.2011.08.008. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 90.Santos RX, Correia SC, Zhu X, et al. Nuclear and mitochondrial DNA oxidation in Alzheimer’s disease. Free radical research. 2012;46:565–576. doi: 10.3109/10715762.2011.648188. [DOI] [PubMed] [Google Scholar]
- 91.Wang J, Xiong S, Xie C, et al. Increased oxidative damage in nuclear and mitochondrial DNA in Alzheimer’s disease. Journal of neurochemistry. 2005;93:953–962. doi: 10.1111/j.1471-4159.2005.03053.x. [DOI] [PubMed] [Google Scholar]
- 92.Cabelof DC, Patel HV, Chen Q, et al. Mutational spectrum at GATA1 provides insights into mutagenesis and leukemogenesis in Down syndrome. Blood. 2009;114:2753–2763. doi: 10.1182/blood-2008-11-190330. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 93.Busciglio J, Yankner BA. Apoptosis and increased generation of reactive oxygen species in Down’s syndrome neurons in vitro. Nature. 1995;378:776–779. doi: 10.1038/378776a0. [DOI] [PubMed] [Google Scholar]
- 94.Sykora P, Misiak M, Wang Y, et al. DNA polymerase beta deficiency leads to neurodegeneration and exacerbates Alzheimer disease phenotypes. Nucleic acids research. 2014 doi: 10.1093/nar/gku1356. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 95.Guzman JN, Sanchez-Padilla J, Wokosin D, et al. Oxidant stress evoked by pacemaking in dopaminergic neurons is attenuated by DJ-1. Nature. 2010;468:696–700. doi: 10.1038/nature09536. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 96.Schapira AH, Cooper JM, Dexter D, et al. Mitochondrial complex I deficiency in Parkinson’s disease. Lancet. 1989;1:1269. doi: 10.1016/s0140-6736(89)92366-0. [DOI] [PubMed] [Google Scholar]
- 97.Bender A, Krishnan KJ, Morris CM, et al. High levels of mitochondrial DNA deletions in substantia nigra neurons in aging and Parkinson disease. Nature genetics. 2006;38:515–517. doi: 10.1038/ng1769. [DOI] [PubMed] [Google Scholar]
- 98.Simon DK, Lin MT, Zheng L, et al. Somatic mitochondrial DNA mutations in cortex and substantia nigra in aging and Parkinson’s disease. Neurobiology of aging. 2004;25:71–81. doi: 10.1016/s0197-4580(03)00037-x. [DOI] [PubMed] [Google Scholar]
- 99.Sanders LH, McCoy J, Hu X, et al. Mitochondrial DNA damage: molecular marker of vulnerable nigral neurons in Parkinson’s disease. Neurobiology of disease. 2014;70:214–223. doi: 10.1016/j.nbd.2014.06.014. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 100.Lin MT, Cantuti-Castelvetri I, Zheng K, et al. Somatic mitochondrial DNA mutations in early Parkinson and incidental Lewy body disease. Annals of neurology. 2012;71:850–854. doi: 10.1002/ana.23568. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 101.Duyao MP, Auerbach AB, Ryan A, et al. Inactivation of the mouse Huntington’s disease gene homolog Hdh. Science. 1995;269:407–410. doi: 10.1126/science.7618107. [DOI] [PubMed] [Google Scholar]
- 102.McMurray CT. Mechanisms of trinucleotide repeat instability during human development. Nature reviews. Genetics. 2010;11:786–799. doi: 10.1038/nrg2828. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 103.Kovtun IV, Liu Y, Bjoras M, et al. OGG1 initiates age-dependent CAG trinucleotide expansion in somatic cells. Nature. 2007;447:447–452. doi: 10.1038/nature05778. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 104.Damiano M, Galvan L, Deglon N, Brouillet E. Mitochondria in Huntington’s disease. Biochimica et biophysica acta. 2010;1802:52–61. doi: 10.1016/j.bbadis.2009.07.012. [DOI] [PubMed] [Google Scholar]
- 105.Yano H, Baranov SV, Baranova OV, et al. Inhibition of mitochondrial protein import by mutant huntingtin. Nature neuroscience. 2014;17:822–831. doi: 10.1038/nn.3721. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 106.Mochel F, Haller RG. Energy deficit in Huntington disease: why it matters. The Journal of clinical investigation. 2011;121:493–499. doi: 10.1172/JCI45691. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 107.Siddiqui A, Rivera-Sanchez S, del Castro MR, et al. Mitochondrial DNA damage is associated with reduced mitochondrial bioenergetics in Huntington’s disease. Free radical biology & medicine. 2012;53:1478–1488. doi: 10.1016/j.freeradbiomed.2012.06.008. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 108.Stack EC, Matson WR, Ferrante RJ. Evidence of oxidant damage in Huntington’s disease: translational strategies using antioxidants. Annals of the New York Academy of Sciences. 2008;1147:79–92. doi: 10.1196/annals.1427.008. [DOI] [PubMed] [Google Scholar]
- 109.Costa V, Scorrano L. Shaping the role of mitochondria in the pathogenesis of Huntington’s disease. The EMBO journal. 2012;31:1853–1864. doi: 10.1038/emboj.2012.65. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 110.Ayala-Pena S. Role of oxidative DNA damage in mitochondrial dysfunction and Huntington’s disease pathogenesis. Free radical biology & medicine. 2013;62:102–110. doi: 10.1016/j.freeradbiomed.2013.04.017. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 111.Xun Z, Rivera-Sanchez S, Ayala-Pena S, et al. Targeting of XJB-5-131 to mitochondria suppresses oxidative DNA damage and motor decline in a mouse model of Huntington’s disease. Cell reports. 2012;2:1137–1142. doi: 10.1016/j.celrep.2012.10.001. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 112.Ebrahimkhani MR, Daneshmand A, Mazumder A, et al. Aag-initiated base excision repair promotes ischemia reperfusion injury in liver, brain, and kidney. Proceedings of the National Academy of Sciences of the United States of America. 2014;111:E4878–4886. doi: 10.1073/pnas.1413582111. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 113.Chen D, Minami M, Henshall DC, et al. Upregulation of mitochondrial base-excision repair capability within rat brain after brief ischemia. Journal of cerebral blood flow and metabolism : official journal of the International Society of Cerebral Blood Flow and Metabolism. 2003;23:88–98. doi: 10.1097/01.WCB.0000039286.37737.19. [DOI] [PubMed] [Google Scholar]
- 114.Lan J, Li W, Zhang F, et al. Inducible repair of oxidative DNA lesions in the rat brain after transient focal ischemia and reperfusion. Journal of cerebral blood flow and metabolism : official journal of the International Society of Cerebral Blood Flow and Metabolism. 2003;23:1324–1339. doi: 10.1097/01.WCB.0000091540.60196.F2. [DOI] [PubMed] [Google Scholar]
- 115.Tsuruya K, Furuichi M, Tominaga Y, et al. Accumulation of 8-oxoguanine in the cellular DNA and the alteration of the OGG1 expression during ischemia-reperfusion injury in the rat kidney. DNA repair. 2003;2:211–229. doi: 10.1016/s1568-7864(02)00214-8. [DOI] [PubMed] [Google Scholar]
- 116.Endres M, Biniszkiewicz D, Sobol RW, et al. Increased postischemic brain injury in mice deficient in uracil-DNA glycosylase. The Journal of clinical investigation. 2004;113:1711–1721. doi: 10.1172/JCI20926. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 117.Liu D, Croteau DL, Souza-Pinto N, et al. Evidence that OGG1 glycosylase protects neurons against oxidative DNA damage and cell death under ischemic conditions. Journal of cerebral blood flow and metabolism : official journal of the International Society of Cerebral Blood Flow and Metabolism. 2011;31:680–692. doi: 10.1038/jcbfm.2010.147. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 118.Sejersted Y, Hildrestrand GA, Kunke D, et al. Endonuclease VIII-like 3 (Neil3) DNA glycosylase promotes neurogenesis induced by hypoxia-ischemia. Proceedings of the National Academy of Sciences of the United States of America. 2011;108:18802–18807. doi: 10.1073/pnas.1106880108. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 119.Canugovi C, Yoon JS, Feldman NH, et al. Endonuclease VIII-like 1 (NEIL1) promotes short-term spatial memory retention and protects from ischemic stroke-induced brain dysfunction and death in mice. Proceedings of the National Academy of Sciences of the United States of America. 2012;109:14948–14953. doi: 10.1073/pnas.1204156109. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 120.Marchesi VT. Alzheimer’s disease and CADASIL are heritable, adult-onset dementias that both involve damaged small blood vessels. Cellular and molecular life sciences : CMLS. 2014;71:949–955. doi: 10.1007/s00018-013-1542-7. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 121.Querfurth HW, LaFerla FM. Alzheimer’s disease. The New England journal of medicine. 2010;362:329–344. doi: 10.1056/NEJMra0909142. [DOI] [PubMed] [Google Scholar]
- 122.Canugovi C, Misiak M, Scheibye-Knudsen M, et al. Loss of NEIL1 causes defects in olfactory function in mice. Neurobiology of aging. 2014 doi: 10.1016/j.neurobiolaging.2014.09.026. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 123.Devanand DP, Lee S, Manly J, et al. Olfactory deficits predict cognitive decline and Alzheimer dementia in an urban community. Neurology. 2014 doi: 10.1212/WNL.0000000000001132. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 124.Israel MA, Yuan SH, Bardy C, et al. Probing sporadic and familial Alzheimer’s disease using induced pluripotent stem cells. Nature. 2012;482:216–220. doi: 10.1038/nature10821. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 125.Lu T, Pan Y, Kao SY, et al. Gene regulation and DNA damage in the ageing human brain. Nature. 2004;429:883–891. doi: 10.1038/nature02661. [DOI] [PubMed] [Google Scholar]
- 126.Horton JK, Joyce-Gray DF, Pachkowski BF, et al. Hypersensitivity of DNA polymerase beta null mouse fibroblasts reflects accumulation of cytotoxic repair intermediates from site-specific alkyl DNA lesions. DNA repair. 2003;2:27–48. doi: 10.1016/s1568-7864(02)00184-2. [DOI] [PubMed] [Google Scholar]
- 127.Harrison JF, Hollensworth SB, Spitz DR, et al. Oxidative stress-induced apoptosis in neurons correlates with mitochondrial DNA base excision repair pathway imbalance. Nucleic acids research. 2005;33:4660–4671. doi: 10.1093/nar/gki759. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 128.Viswanathan A, You HJ, Doetsch PW. Phenotypic change caused by transcriptional bypass of uracil in nondividing cells. Science. 1999;284:159–162. doi: 10.1126/science.284.5411.159. [DOI] [PubMed] [Google Scholar]
- 129.Galashevskaya A, Sarno A, Vagbo CB, et al. A robust, sensitive assay for genomic uracil determination by LC/MS/MS reveals lower levels than previously reported. DNA repair. 2013;12:699–706. doi: 10.1016/j.dnarep.2013.05.002. [DOI] [PubMed] [Google Scholar]
- 130.Scheibye-Knudsen M, Ramamoorthy M, Sykora P, et al. Cockayne syndrome group B protein prevents the accumulation of damaged mitochondria by promoting mitochondrial autophagy. The Journal of experimental medicine. 2012;209:855–869. doi: 10.1084/jem.20111721. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 131.Cummings JL, Morstorf T, Zhong K. Alzheimer’s disease drug-development pipeline: few candidates, frequent failures. Alzheimer’s research & therapy. 2014;6:37. doi: 10.1186/alzrt269. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 132.Frost B, Gotz J, Feany MB. Connecting the dots between tau dysfunction and neurodegeneration. Trends in cell biology. 2015;25:46–53. doi: 10.1016/j.tcb.2014.07.005. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 133.Fu D, Calvo JA, Samson LD. Balancing repair and tolerance of DNA damage caused by alkylating agents. Nature reviews. Cancer. 2012;12:104–120. doi: 10.1038/nrc3185. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 134.Parsons JL, Tait PS, Finch D, et al. CHIP-mediated degradation and DNA damage-dependent stabilization regulate base excision repair proteins. Molecular cell. 2008;29:477–487. doi: 10.1016/j.molcel.2007.12.027. [DOI] [PubMed] [Google Scholar]
- 135.Parsons JL, Dianova, Khoronenkova SV, et al. USP47 is a deubiquitylating enzyme that regulates base excision repair by controlling steady-state levels of DNA polymerase beta. Molecular cell. 2011;41:609–615. doi: 10.1016/j.molcel.2011.02.016. [DOI] [PubMed] [Google Scholar]
- 136.Hegre SA, Saetrom P, Aas PA, et al. Multiple microRNAs may regulate the DNA repair enzyme uracil-DNA glycosylase. DNA repair. 2013;12:80–86. doi: 10.1016/j.dnarep.2012.10.007. [DOI] [PubMed] [Google Scholar]

